Unraveling The Hook Effect

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

This is an open access article published under a Creative Commons Non-Commercial No

Derivative Works (CC-BY-NC-ND) Attribution License, which permits copying and


redistribution of the article, and creation of adaptations, all for non-commercial purposes.

pubs.acs.org/ac Article

Unraveling the Hook Effect: A Comprehensive Study of High


Antigen Concentration Effects in Sandwich Lateral Flow
Immunoassays
Georgina M. S. Ross, Daniel Filippini, Michel W. F. Nielen, and Gert IJ. Salentijn*
Cite This: Anal. Chem. 2020, 92, 15587−15595 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Sandwich lateral flow immunoassays (LFIAs) are


limited at high antigen concentrations by the hook effect, leading
to a contradictory decrease in the test line (T) intensity and false-
Downloaded via 186.249.229.86 on September 6, 2022 at 17:15:47 (UTC).

negative results. The hook effect is mainly associated with the loss
of T, and research focuses on minimizing this effect. Nevertheless,
the control line (C) intensity is also affected at higher analyte
concentrations, undesirably influencing the T/C ratio in LFIA
readers. The main aim of this work is to identify and understand
these high antigen concentration effects in order to develop
ubiquitous strategies to interpret and mitigate such effects. Four
complementary experiments were performed: performance assess-
ment of three different allergen LFIAs (two for hazelnut, one for
peanut) over 0.075−3500 ppm, LFIAs with C only, surface
plasmon resonance (SPR) binding experiments on the immobilized control antibody, and smartphone video recording of LFIAs
during their development. As antigen concentrations increase, the C signal decreases before the T signal does, suggesting that
distinct mechanisms underlie these intensity reductions. Reduced binding at the C occurred even in the absence of T, so the upfront
T does not explain the loss of C. SPR confirmed that the C antibody favors binding with free labeled antibody compared with a
labeled antibody−analyte complex, indicating that in antigen excess, binding is reduced at C before T. Finally, a smartphone-based
video method was developed for dynamically monitoring the LFIA development in real time to distinguish between different
concentration-dependent effects. Digitally analyzing the data allows clear differentiation of highly positive samples and false-negative
samples and can indicate whether the LFIA is in the dynamic working range or at critically high concentrations. The aim of this work
is to identify and understand such high antigen concentration effects in order to develop ubiquitous strategies to interpret and
mitigate such effects.

ateral flow immunoassays (LFIAs) have revolutionized


L consumer diagnostics, translating laboratory-based immu-
noassays into affordable and accessible home testing devices.1
intensity.3,5−7 This disappearance of T is known as the hook
effect.8
In LFIA, in addition to the capture and detection mAbs, a
Sandwich-format LFIAs utilize two bivalent monoclonal secondary species-specific antibody, capable of binding the
antibodies (mAbs) to capture and detect large multivalent labeled detection mAb, is immobilized as a control line (C).9
targets, such as allergens.2 In microplate-based immunoassays, The C informs the user that the test is valid, yielding a signal
capture mAbs are directly immobilized onto the solid support regardless of the presence of an antigen. When analyzing LFIAs
of the microwell;3 in LFIA, the capture mAb is shaped into a with a digital optical reader such as a smartphone,10 the ever-
test line (T). A labeled secondary mAb, which generally binds present C can be used to normalize the T, to correct for
to a different or repeating epitope on the antigen, forms a experimental variables (T/C).11 The use of T/C thus assumes
sandwich complex with the antigen and the capture antibody. that the C intensity is constant. However, it has been observed
The label yields a measurable, often optical signal. In a that an increasing antigen concentration also leads to a
sandwich LFIA’s working range, the T signal increases with an
increase in the target antigen concentration; the naked eye can Received: September 3, 2020
qualitatively read this. However, researchers have known since Accepted: November 2, 2020
19744 that an excess antigen concentration leads to saturation Published: November 13, 2020
of available binding sites on the bivalent capture and detector
mAbs, preventing the formation of a sandwich complex in the
T area, which in turn leads to a paradoxical loss of T signal

© 2020 American Chemical Society https://dx.doi.org/10.1021/acs.analchem.0c03740


15587 Anal. Chem. 2020, 92, 15587−15595
Analytical Chemistry pubs.acs.org/ac Article

decrease in the C intensity, while the T intensity still Reference Material Preparation. Standardized certified
increases.12−14 The loss of C compromises the reliability of reference materials and standard solutions for food allergens
the T/C at high concentrations and yet remains to be fully are not currently available, and therefore, the total protein
understood. In the literature, various concentration-dependent extracts used here required in-house preparation.27 The
effects are described under the hook-effect definition. See the procedure for the protein extraction for generating total
Supporting Information, Table S1, for a review of the peanut protein (TPP) and total hazelnut protein (THP) has
definitions given for the hook effect and the observed effects been described previously.25,26 Fresh aliquots were defrosted
in the literature (1974−2020). Despite the qualifying on the day of the experiment, and the protein content was
characteristic of the hook-effect being a false negative result checked with the NanoDrop (ND 3300, Isogen Life Sciences,
(i.e., the absence of T), the definition is frequently also used to De Meern; The Netherlands) before use.
describe effects causing the loss of the C. LFIA Readout. A qualitative assessment of LFIAs was made
There are numerous mitigation strategies to cope with high- by reading the developed signal with the naked eye;
concentration effects in sandwich immunoassays: quantitative readings were performed by smartphone detec-
The most apparent method is testing the sample, both tion.28 A custom 3D-printed smartphone holder was used to
undiluted and diluted.15 If the diluted sample gives a stronger shield up to three LFIAs from ambient light during the optical
T response than the undiluted sample, the undiluted result can measurements (See Figures 1, S1 and S2). A smartphone
be considered as “hooked”.16 Dilutions allow adjustment of the
dynamic working range of an LFIA but also require additional
sample preparation, time, and material costs.17 Conversely,
changing the physical layout of the assay can prevent high-
concentration effects,2,18 essentially allowing for decoupled
reagent delivery.19 Still, the separation of reagent flow is crucial
to prevent premature mixing of the labeled antibody and
analyte.20 Alternatively, high-concentration effects can be
minimized by optimizing reagents, for example, by supple-
menting the immunoassay with one or more additional target
lines.6,21,22 Differentiation between artificially low (“hooked”)
and truly low concentration samples is possible by the real-
time monitoring of T and C development14 and can allow for
LFIA dynamic ranges to be expanded by orders of Figure 1. Photographs of the 3D-printed smartphone holder for
magnitude.23 recording LFIA signal development under controlled lighting: (A)
In this work, the aim was to first unravel the hook effect by open side view to show where LFIAs and microwells are inserted; (B)
comprehensively elucidating how extreme antigen concen- closed side view with the LFIAs inserted and smartphone attached;
trations influence the LFIA test line and control line and (C) head-on view of LFIA signal development on the phone
development over time in three different allergen LFIAs. The screen.
identification and understanding of how high-antigen concen-
trations influence LFIA signal development is crucial for any (Google Pixel 2 XL, Google, Mountain View, CA; USA) was
developed sandwich LFIAs and will lead to ways to mitigate used to record images and videos of developing LFIAs. The
such effects, such as the simplified dynamic smartphone-based smartphone was attached to the frame of the holder for
method presented here, ultimately leading to more reliable support during the dynamic measurements (locked exposure,
testing. fixed focus, and controlled illumination) at 30 frames per

■ EXPERIMENTAL SECTION
Three allergen LFIAs were developed for detecting peanut
second (fps) using OpenCamera (v1.47.3). Adapter (v2.1.6)
converted the videos into images of 1 fps. ImageJ29 was used to
split the images into their RGB (red, green, and blue) color
(PA) and hazelnut (HA1 + HA2). Each assay has a different channels. Blue channel intensity values for the T and C were
sandwich pair of mAbs for their capture (T) and detector subtracted from a background reading; the resulting corrected
carbon nanoparticle-labeled-antibody (CNP-mAb), selected blue channel pixel intensity (cBCPI) increased with increasing
for their differences in sensitivity as observed in prior line intensity.
works.24−26 All assays used goat anti-mouse (GAMaB) IgG Influence of Antigen Concentration on LFIA Signal
in PBS (pH 7.6; 1.2 mg/mL; AffiniPure F(ab’)2 Fragment) at Development. All three LFIAs were tested in a concentration
the C line (Jackson Immunoresearch Laboratories Inc., Sanbio, range spanning 5 orders of magnitude. The LFIAs were
Uden; The Netherlands) and were developed on nitrocellulose inserted into microwells containing 99 μL of antigen in RB
membranes (140 CN; Unisart, Sartorius, Gottingen, Germany) (0.0075−3500 ppm of TPP or THP) and 1 μL of CNP-mAb.
overlaid with an absorbent pad (Whatman, GE Healthcare, Here, the endpoint images of T and C signals were used for
Eindhoven, The Netherlands) and secured with a plastic calculating T/C; LFIAs were left to develop for 40 min before
backing (G and L, San Jose, CA, USA); see Supporting the results were recorded using a smartphone.
Information Protocol S1 and S2 for full details on CNP-mAb Influence of Antigen Concentration on Signal
labeling and LFIA preparation details. For the surface plasmon Development in C-Only LFIAs. The LFIAs were tested
resonance (SPR) biosensor assay, an amine coupling kit, pH alongside an LFIA with only a GAMaB C, at concentrations
scouting kit, HBS-EP buffer, and CM5 sensor chips were shown to diminish C intensity (5, 10, 25, and 50 ppm) (n = 2).
purchased from GE Healthcare (Uppsala, Sweden); see Additionally, the original LFIAs were run in a blank as a
Supporting Information Protocol S3 for further details on negative control. The LFIAs were developed in a microwell
the immobilization procedure. containing 99 μL of antigen in RB (50 ppm TPP, 5 or 50 ppm
15588 https://dx.doi.org/10.1021/acs.analchem.0c03740
Anal. Chem. 2020, 92, 15587−15595
Analytical Chemistry pubs.acs.org/ac Article

Figure 2. Extended calibration range of the three LFIAs, peanut assay [PA], hazelnut assay 1 [HA1], and hazelnut assay 2 [HA2] in increasing
concentration of the analyte (TPP or THP) spiked into RB (0.0075−3500 ppm). (A,D,G) Photographs after 40 min, (B,E,H) test and control
signal expressed in cBCPI, and (C,F,I) test line divided by the control line (T/C ratio). Error bars represent the standard deviation (n = 3).

15589 https://dx.doi.org/10.1021/acs.analchem.0c03740
Anal. Chem. 2020, 92, 15587−15595
Analytical Chemistry pubs.acs.org/ac Article

THP) and 1 μL of CNP-mAb; the LFIAs were developed for negatives. This hook effect is unsurprising because, at extreme
30 min in a 3D-printed smartphone holder (Figure 1). antigen-excess, T is rapidly saturated by an accumulation of
Sequential and Premixed Antigen-Binding Studied unlabeled antigen, while the remaining mobile, excess antigen
by SPR. Biosensor chip immobilization was performed per the binds with CNP-mAb without being captured on T.2
manufacturer’s instructions; see Supporting Information Comparably, the C signal arises because of the binding of
Protocol S3 for further details. Using SPR, the influence of CNP-mAb. Therefore, in a blank sample (0 ppm), a clear C is
increased analyte concentration on the binding characteristics seen. At low concentrations, C captures free CNP-mAbs,
of free CNP-mAb (and subsequently introduced analyte) or yielding an intense signal, as observed visually (Figure 2A,D,G)
premixed immunocomplex (CNP-mAb-analyte) toward im- and numerically (Figure 2B,E,H). In this range, C intensity
mobilized GAMaB was evaluated (n = 3). For sequential increases slightly with increasing antigen concentration,
measurements, 10 μL of CNP-mAb (diluted 1:99 in HBS-EP) possibly because of the binding of multiple CNP-mAbs to
was injected across the GAMaB surface (flow rate of 25 μL/ the same multivalent allergenic protein, resulting in increased
min). After binding, 10 μL of analyte solution was injected. For signal intensity.31 At higher concentrations [roughly above 10
premix experiments, 10 μL of CNP-mAb and analyte solution ppm (PA), 0.5 ppm (HA1), and 1 ppm (HA2)], the C
was injected at 25 μL/min. These experiments were performed intensity decreases, while the T still becomes more intense.
at concentrations of 0.1−3000 ppm. For regenerating the This is reflected in the endpoint T/C metric (Figure 2C,F,I).
surface, the flow was adjusted (50 μL/min) and 5 μL of 25 T/C increases along with T and rises further even at
mM NaOH was injected, returning the signal to baseline. concentrations causing C to decrease. This increase in T/C
Dynamic Monitoring of LFIA Signal Development. widens the LFIA’s linear dynamic working range even when C
The LFIAs were placed in microwells containing 99 μL of is affected by the concentration. Only after the hook-effect has
antigen in RB (0.0075−3500 ppm of TPP or THP) and 1 μL occurred, leading to a decrease in T, does the T/C drastically
of CNP-mAb, inserted into the smartphone holder, and were drop. This trend is consistent across all three assays, despite
dynamically recorded for 30 min. The T/C was acquired at set them detecting distinct antigens and using different mAb
time points (5−30 min) by selecting frames from the video. sandwich pairs with diverse sensitivities and kinetics;24,25 as
This was done for different concentrations [(PA; 0.015−3000 well as from the body of literature describing high
ppm), (HA1: 0.015−3000 ppm), and (HA2: 0.015−2000 concentration effects in LFIA (see Supporting Information
ppm)]. Alternatively, the videos were imported into custom Table S12,12,14). A decrease in C must be caused by a
python scripts (Python 3.8) for automated data analysis. In an reduction in CNP-mAb binding, which could potentially be
early video frame, regions of interest (ROI) were positioned due to preventing the arrival of CNP-mAbs at the C by the T
over the T and C, and as a background reading. Data (investigated in the C-only section below) or decreasing the
evaluation consisted of averaging the blue pixel intensity in the avidity of the CNP-mAbs to bind at those sites, which is
ROIs across the entire video duration at 1 second data points. assessed by the SPR experiments.32
The generated responses were exported in a comma-separated Interestingly, as the concentration further increases [above
value (.csv) format for easy importing into spreadsheet 100 ppm (PA) and 250 ppm (HA1 and HA2)], the C intensity
programs. A second complementary python script corrected partially reappears. At extreme concentrations, multivalent
the time response to assess LFIA signal development. proteins have a propensity to aggregate, potentially masking

■ RESULTS AND DISCUSSION


Influence of Antigen Concentration on LFIA Signal
their epitopes.33 Moreover, these antigen concentrations
probably hinder higher-order immunocomplex formation
because there is insufficient CNP-mAb available for binding
Development. As most reported LFIAs are tested within a with the abundant antigen in larger complexes.12,34
limited range and read after up to 20 min, high concentration Influence of Antigen Concentration on Signal
effects are not well documented, but it is known that excess Development in C-Only LFIAs. In the extended calibration
antigen concentrations can influence the signal development range, specific concentrations caused the appearance of an
time. Here, three different LFIAs were tested in a intense T and a comparably diminished C, affecting the T/C,
concentration range spanning 0.0075−3500 ppm (Figure 2) as was observed in Figure 2C,F,I. Possibly, at these moderately
and developed over 40 min before their endpoint image was high concentrations, the antigen, which is bound in the
recorded using a smartphone. immunocomplex with CNP-mAb, binds mostly at the T, thus
In all three assays, the T follows the same pattern across the limiting the amount of CNP-mAb that can reach the C and can
concentration range, as can be observed visually (Figure interact with GAMaB.
2A,D,G) and numerically (Figure 2B,E,H). The T signal Figure 3A−C shows the three assays tested at concentrations
depends on the capture of antigen followed by binding of observed to affect C development; Figure 3D shows the
CNP-mAb, or on the binding of already complexed [CNP- cBCPIs (n = 2) of the C from the three variations of the assay,
mAb-analyte] to the immobilized antibody. At 0 ppm, there is for all the three assays. Both in the presence and absence of a
no analyte present; therefore, no T signal develops. As the T, the C never reaches the full intensity it would reach in a
analyte concentration increases [up to 100 ppm (PA), 10 ppm blank. Interestingly, the signal for the C in the regular LFIA
(HA1), and 10 ppm (HA2)], the T intensity increases by [green (PA), blue (HA1), and orange (HA2)] and the C-only
capturing more analyte and correspondingly more CNP-mAb LFIA [checked green (PA), checked blue (HA1), and checked
or by capturing larger, higher-order immunocomplexes. These orange (HA2)] are of similar intensity. However, 3D shows
complexes form when multivalent antigens, such as hazelnuts that the intensity of C in the C + T LFIAs is often less than in
and peanuts, with numerous identical or distinct epitopes,30 the C-only LFIAs, indicating that some binding of [CNP-mAb-
bind several CNP-mAb particles, leading to the formation of an analyte] at T could have a minor contribution to the reduction
intense T at high concentrations. Beyond these concentrations, of binding at C. However, there is a more substantial C
the T intensity instead starts decreasing, producing false intensity difference between the blank and antigen solutions
15590 https://dx.doi.org/10.1021/acs.analchem.0c03740
Anal. Chem. 2020, 92, 15587−15595
Analytical Chemistry pubs.acs.org/ac Article

Figure 3. Control line-only LFIAs. Control signal development in


LFIAs with a control line and a test line (C + T) and LFIAs with only
a control line (C). (A) Peanut assay (PA) C + T signal development
in a blank (B), C + T and C in 50 ppm TPP. (B) Hazelnut assay 1
(HA1) C + T and C signal development in B and in 5 ppm THP, (C)
hazelnut assay 2 (HA2) C + T and C signal development in B and in
50 ppm THP, and (D) signal intensity in B, C + T and C across all
three LFIAs as a cBCPI.

for each assay, at all tested concentrations. This emphasizes


that a reduction of binding at C must be due to the increased
antigen concentrations, causing the CNP-mAb to have a
decreased avidity for C.
Sequential and Premixed Antigen-Binding Studied
by SPR. Here, using SPR, we set out to elucidate whether
increased antigen concentration hinders the labeled [CNP-
mAb-analyte] immunocomplex’ ability to bind with the
GAMaB (premix) compared with whether increased antigen
concentration affects the binding of CNP-mAb decoupled
from the analyte. SPR typically is label-free,24 but antibody
labeling can alter essential binding characteristics.35,36 In SPR,
response units (RUs) are generated by the total amount of
material captured at the surface, compared with LFIA, where
the signal is made up solely from the binding of CNP-mAb to
C. Moreover, it is also important to note that these assays take
place on a very different time scale (i.e., LFIA; 40 min vs SPR;
40 s).
The RU’s reproducibly increased in all three assays following
the injection of CNP-mAb, as can be seen in Figure 4A−C, Figure 4. SPR responses showing binding to goat anti-mouse
where the black bar represents the free CNP-mAb binding with antibody of (A) total peanut protein [PA], (B) total hazelnut protein
GAMaB. Adjusting the concentration of the injected analyte in 1 [HA1], and (C) total hazelnut protein 2 [HA2] tested by
the second step again leads to a reproducible RU increase in all sequentially injecting the carbon nanoparticle-labeled antibody
assays, as can be seen by the increasing colored bar [green (CNP-mAb; black) followed by antigen [green (PA), blue (HA1),
(PA), blue (HA1), orange (HA2)]. Because these multivalent and orange (HA2)] compared against premixed CNP-mAb + antigen
antigens bind to the captured CNP-mAb, thereby increasing [checked green (PA), checked blue (HA1), and checked orange
the total mass of material bound to the chip surface, this (HA2])] (n = 3). Standard deviation is expressed as error bars (n =
increase in RU is unsurprising. 3).
Contrastingly, when simultaneously injecting CNP-mAb +
antigen (premix), the assays behaved differently [checked
green (PA), checked blue (HA1), and checked orange (HA2)] complex.2 This explains why high antigen concentrations cause
(see Supporting Information Figure S3 for an example of a reduction in binding toward GAMaB in both LFIA and SPR.
sequential and premixed TPP sensorgrams). Crucially, the The C signal is consistently reduced at a lower concentration
signal intensity in the premix approach is consistently lower than the T signal in LFIA. As soon as an assay is in antigen
versus the sequential approach at all tested antigen excess, a prerequisite to enter the hook-range, higher-order
concentrations. In excess analyte concentrations, [CNP-mAb- immunocomplexes would already have formed in the solution.
analyte] quickly forms, depleting the free CNP-mAb, which Formation of such complexes would deplete the amount of
would otherwise interact with C with higher avidity than the free CNP-mAb available for binding with C.
15591 https://dx.doi.org/10.1021/acs.analchem.0c03740
Anal. Chem. 2020, 92, 15587−15595
Analytical Chemistry pubs.acs.org/ac Article

Figure 5. Dynamic smartphone monitoring of signal development for total peanut protein (PA), total hazelnut protein 1 (HA1), and total hazelnut
protein 2 (HA2) LFIAs. Control line signal development (A) [PA], (B) [HA1], and (C) [HA2]; test line signal development (D) [PA], (E)
[HA1], and (F) [HA2]; and T/C ratio development (G) [PA], (H) [HA1], and (I) [HA2] over 30 min at different concentrations [0.015−5 ppm
(PA)] [0.015−0.125 ppm (HA1 and 2)].

Therefore, for the hook effect at T to occur, the reduction in C in the corresponding LFIAs (Figures 2 and 3).
concentration effect on the C must have already taken place. However, at higher concentrations, both HA assays instead
The response in both [HA] assays is different from the [PA] increase in RU; these concentrations likely cause protein
assay, although this is not surprising, considering that all the aggregation and have a relative scarcity of CNP-mAbs
assays use different antibodies with varied sensitivities and compared to the overabundant antigen, inhibiting the
detect distinct analytes. In the [PA], the total binding formation of higher-order complexes.37
decreases as antigen concentration increases, which is Dynamic Monitoring of LFIA Signal Development.
consistent with the formation of a higher number of The SPR data suggest that the binding of immunocomplexes
immunocomplexes with reduced avidity to GAMaB. Pre- differs from the binding of free CNP-mAb, resulting in
viously, using SPR, Liang2 observed a concentration-depend- variations of the C intensity across broad concentration ranges.
ent decrease in labeled immunocomplex binding toward T, but The usefulness of T/C for normalizing sandwich LFIA results
did not study how concentration influences complex binding at is impeded across this range when considering only the
C. Here, it is clear that there is a decrease in binding at endpoint analysis. However, recent research suggests that
GAMaB with increasing TPP concentrations, seemingly additional information is available by monitoring the develop-
harmonious with a growing number of complexes. In both ment of the T/C over the entire assay duration.14 Figure 5
[PA] formats, a decrease in C signal is observed at high shows the signal development of the C (Figure 5A−C), T
concentrations, but this stabilizes going up to 3000 ppm. In the (Figure 5D−F), and T/C (Figure 5G−I) of the LFIAs at a
LFIA, one additional higher concentration (3500 ppm) was range of TPP or THP concentrations [0.015−3000 ppm (PA)]
also tested, and at this concentration, C partially recovered. [0.015−3000 ppm (HA1) and 0.015−2000 ppm (HA2)] (see
Contrastingly, in both [HA] LFIAs, we see a reappearance of Supporting Information Figure S4 for snapshots from the video
C at much lower concentrations than with [PA], which is recording of developing LFIAs).
logical, considering that the [HA] LFIAs are more sensitive Within the assay working range [0.015−5 ppm (PA)]
than the [PA]. Likewise, both premixed HAs initially decrease [0.015−0.125 ppm (HA1 and 2)], the C always develops faster
in RU with increasing concentration (<100 ppm), consistent than or at the same time as the T, resulting in a low, stable T/
with the formation of higher-order complexes with reduced C over time. With increasing concentrations [10−250 ppm
avidity for C; these concentrations are also observed to cause a (PA)], [2.5−100 ppm (HA1)], and [2.5−400 ppm (HA2)],
15592 https://dx.doi.org/10.1021/acs.analchem.0c03740
Anal. Chem. 2020, 92, 15587−15595
Analytical Chemistry pubs.acs.org/ac Article

Table 1. Summary of Results of Concentration-Dependent Effects Studied in This Work


experiment purpose result conclusions
influence of concentration on investigate T and C and T/C development C increases at low [conc] hook effect is reproducible
LFIA signal development (0.0075−3500 ppm)
T increases at high [conc], at extreme
[conc] T is lost (hook effect)
C decreases at a lower [conc] than T C decreases because of reduced
CNP-mAb binding or due to T
depleting CNP-mAb
C partially recovers at extreme [conc]
influence of concentration on determine how C signal develops without a T high [conc] prevents C from reaching high [conc] negatively affects C
signal development in C only the same intensity as in a blank
LFIAs
sequential and determine the difference between premix and sequential sequential [PA], [HA1], [HA2]: at high free CNP-mAb has higher avidity to
premixed antigen-binding binding of CNP-mAb and antigen to GAMaB [conc] RUs increase GAMaB than [CNP-mAb-analyte]
studied by SPR (0.1−3000 ppm) complexes
premix [PA]: at high [conc] RUs
decrease
premix [HA1] + [HA2]: at low [conc]
RUs decrease and at high [conc] RUs
increase
dynamic monitoring of LFIA investigate T, C and T/C development over time at dynamic working range: T and C at a dynamic monitoring of T and C can
signal development varying concentrations similar time distinguish high concentration ef-
fect
high [conc]: T increases faster than C;
T/C increases and then decreases
extreme [conc]: no signal for approx.
10 min, then C increases

the speed of binding shifts, with the T developing more quickly eventually starts to develop after 10 min, giving a final low T/
than the C. The T/C time development reflects this with a C.
sharp initial increase (5 min), when mostly T is present, Critically, at these concentrations, the resulting low final T/
followed by a steady decline as the C belatedly develops and C overlaps with the T/C at much lower concentrations [0.015
the signal balances out. Outside of the dynamic working range, ppm (PA), 0.015 and 0.03 ppm (HA1) and (HA2)];
when the T or C development is influenced by antigen misinterpretation of this could lead to reporting of a false
concentration, the final T/C becomes an unreliable metric. negative, as can be seen in Supporting Information Figure S6,
Recently, Rey14 presented a method for monitoring LFIA where the T and C signal development and resulting T/C are
signal development, using a different label (i.e., gold nano- compared for HA1 at 0.015 ppm and 2000 ppm. While the
particles), detecting another analyte (C-reactive protein), and final T/Cs at these concentrations are similar, a false negative
similarly observed an initial rise, followed by a decrease in T/C can be avoided by monitoring when the T and C develop; at
over the assay duration at concentrations where the T develops hook-effect concentrations, no signal develops on either line
before the C. Interestingly, here we observed this trend across for the first 10 min, causing a static T/C during this time.
all three LFIAs, albeit at different concentrations. Unfortu- Comparatively, at low concentrations in the dynamic working
nately, Rey’s14 study primarily looked at T/C’s in a limited range, the C signal develops rapidly and with a high intensity.
antigen range (120−255 μg/mL), and as such only tested By using dynamic video acquisition, real and artificially low
concentrations that caused a delayed and diminished develop- concentration measurements can be differentiated, not only
ment of C rather than the inhibition of T development. A based on signal intensity but additionally based on whether the
recent study23 consolidated Rey’s results by testing a different C or T develops first (Figure 5). Further, it is possible to
antigen (hCG) and modeling and observing T/C development automatically generate these T and C development profiles
at different antigen concentrations (0.5−500 IU/mL) over 10 directly from the smartphone video (see Supporting
min. They found that at low hCG concentrations, C initially Information Figure S7 for automatic profiles for PA compared
develops faster than T, with T/C increasing after C is saturated with manual time-development graphs and Figure S8 for a
and T keeps developing over 10 min. Conversely at high blank assay), using the python script. In addition to dynamic
concentrations T develops rapidly with C developing slowly, intensity measurements, this allows for BCPI correction and
autobackground subtraction. Considering that the results
causing T/C to intially rise and afterwards to steadily decrease
generated by the software correlate with the manually plotted
over the assay duration as T is saturated and C still increases.
time responses, novice users can simply use the automated
which is in line with our own findings.
results rather than carrying out the image analysis, corrections,
Additionally, we found that a delayed and weak C
and background subtraction themselves.


development at higher concentrations also occurs when testing
LFIAs without a T [50 ppm (PA), 25 ppm (HA1), and 50
ppm (HA2)] (Supporting Information Figure S5), where the CONCLUSIONS
C signal only starts developing after 10 min and then only with The experiments, results, and conclusions drawn have been
a diminished intensity. As concluded from the SPR study, summarized in Table 1; a schematic depiction of the
reduction in C intensity is mainly caused by the formation of complementary experiments can be found in Supporting
(higher-order) immunocomplexes, which have reduced avidity Information Figure S9. To unravel the hook effect, we devised
for C. While any T signal is weak, a diminished and delayed C an inexpensive, dynamic, smartphone-based method for
15593 https://dx.doi.org/10.1021/acs.analchem.0c03740
Anal. Chem. 2020, 92, 15587−15595
Analytical Chemistry pubs.acs.org/ac Article

directly identifying the concentration-dependent effects across Authors


three different sandwich LFIAs. We comprehensively eluci- Georgina M. S. Ross − Wageningen Food Safety Research
dated how both the antigen concentration and time influence (WFSR), Wageningen University & Research, Wageningen
signal development and allow us to differentiate between two 6700 AE, The Netherlands; orcid.org/0000-0002-2145-
distinct concentration effects: (1) the reduced development of 2386
a C in the presence of a rapidly and intensely developing T, Daniel Filippini − Optical Devices Laboratory, Division of
which occurs within the first few minutes and (2) the decrease Sensor and Actuator Systems, IFMLinköping University,
in signal development on either line for 10 min, followed by Linköping S58183, Sweden
the development of a C which increases in intensity for the Michel W. F. Nielen − Wageningen Food Safety Research
remaining assay duration. These trends were consistent across (WFSR), Wageningen University & Research, Wageningen
all three assays. 6700 AE, The Netherlands; Laboratory of Organic
Indeed, we discovered for the hook effect to occur on T, the Chemistry, Wageningen University, Wageningen 6708 WE,
concentration effect at C must have already happened. Based The Netherlands; orcid.org/0000-0003-4634-0249
on our findings, we propose that a more appropriate definition Complete contact information is available at:
for the moderate-high concentrations, which lead to the loss of https://pubs.acs.org/10.1021/acs.analchem.0c03740
the C in LFIA, would be the situation of free secondary antibody
depletion. In f ree secondary antibody depletion, higher-order Author Contributions
[CNP-mAb-analyte] immunocomplexes form, hindering the G.M.S.R. conceptualization; methodology; validation; formal
signal development at C by reducing the CNP-mAb avidity for analysis; investigation; data analysis; writing original draft;
binding. As the concentration of analyte increases further, the writingreview and editing; project administration D.F.
assay starts to enter the hook-effect range. At these extreme software; review and editing; M.W.F.N. resources; review
concentrations, the unlabeled analyte rapidly saturates T. We and editing; supervision; funding acquisition; project admin-
have established that while the endpoint T/C is an appropriate istration G.IJ.S. conceptualization; methodology; writing
metric within the dynamic working range, outside of this range review and editing; supervision; project administration.
when the test or control line is falsely diminished, the final T/
C is influenced. While a prediction algorithm is outside of the Funding
scope of this paper, we appreciate that this would be a useful This project has received funding from the European Unionʼs
advancement. Here, the automatically generated qualitative Horizon 2020 research and innovation program under the
binding profiles provide a simplified way for novice users to Marie Sklodowska-Curie grant agreement no. 720325.
monitor concentration effects without performing any image or Notes
data analysis themselves. The authors declare no competing financial interest.
Further, the current system can simultaneously analyze three
LFIAs, making it feasible to include an in-built quality control
LFIA; such a control would be highly relevant where non-
■ REFERENCES
(1) Ross, G. M. S.; Bremer, M. G. E. G.; Nielen, M. W. F. Anal.
experts, such as allergic consumers, perform LFIAs. Ultimately, Bioanal. Chem. 2018, 410, 5353−5371.
the use of dynamic readout provides an inexpensive and direct (2) Liang, T.; Robinson, R.; Houghtaling, J.; Fridley, G.; Ramsey, S.
mechanism for identifying the high-concentration effects in A.; Fu, E. Anal. Chem. 2016, 88, 2311−2320.
LFIA. The digital analysis of dynamic data allows clear (3) Hoofnagle, A. N.; Wener, M. H. J. Immunol. Methods 2009, 347,
differentiation between the highly concentrated samples and 3−11.
low concentration results. We foresee that this method should (4) Miles, L. E. M.; Lipschitz, D. A.; Bieber, C. P.; Cook, J. D. Anal.
Biochem. 1974, 61, 209−224.
have broad applicability for distinguishing false-negative results
(5) Petakov, M. S.; Damjanović, S. S.; Nikolić-Durović, M. M.;
in sandwich LFIAs. Dragojlović, Z. L.; Obradović, S.; Gligorović, M. S.; Simić, M. Ž .;


*
ASSOCIATED CONTENT
sı Supporting Information
Popović, V. P. J. Endocrinol. Invest. 1998, 21, 184−188.
(6) Hu, J.; Choi, J. R.; Wang, S.; Gong, Y.; Feng, S.; Pingguan-
Murphy, B.; Lu, T. J.; Xu, F. Sens. Actuators, B 2017, 243, 484−488.
(7) Ki, H.; Oh, J.; Han, G.-R.; Kim, M.-G. Lab Chip 2020, 20, 844−
The Supporting Information is available free of charge at 851.
https://pubs.acs.org/doi/10.1021/acs.analchem.0c03740. (8) O’Farrell, B. Evolution in Lateral Flow−Based Immunoassay
Systems. In Lateral Flow Immunoassay; Wong, R., Tse, H., Eds.;
Overview of hook-effect definitions; protocols for CNP-
Humana Press: Totowa, NJ, 2009; pp 1−33.
mAb labeling, LFIA preparation and SPR procedures; (9) Kulabhusan, P. K.; Rajwade, J. M.; Sugumar, V.; Taju, G.; Sahul
3D-printed attachments, SPR sensorgrams, photos of Hameed, A. S.; Paknikar, K. M. PLoS One 2017, 12, No. e0169012.
developing LFIAs, automatically generated kinetic (10) Nelis, J. L. D.; Tsagkaris, A. S.; Dillon, M. J.; Hajslova, J.;
profiles, and a schematic overview of paper (PDF) Elliott, C. T. TrAC, Trends Anal. Chem. 2020, 129, 115934.


(11) Foysal, K. H.; Seo, S. E.; Kim, M. J.; Kwon, O. S.; Chong, J. W.
Sensors 2019, 19, 4812.
AUTHOR INFORMATION (12) Posthuma-Trumpie, G. A.; Wichers, J. H.; Koets, M.;
Corresponding Author Berendsen, L. B. J. M.; van Amerongen, A. Anal. Bioanal. Chem.
2012, 402, 593−600.
Gert IJ. Salentijn − Wageningen Food Safety Research (13) Pilavaki, E.; Demosthenous, A. Sensors 2017, 17, 2673.
(WFSR), Wageningen University & Research, Wageningen (14) Rey, E. G.; O’Dell, D.; Mehta, S.; Erickson, D. Anal. Chem.
6700 AE, The Netherlands; Laboratory of Organic 2017, 89, 5095−5100.
Chemistry, Wageningen University, Wageningen 6708 WE, (15) Ruff, L. E.; Sapre, A. A.; Plaut, J. S.; De Maere, E.; Mortier, C.;
The Netherlands; orcid.org/0000-0002-2870-9084; Nguyen, V.; Separa, K.; Vandenbogaerde, S.; Vandewalle, L.; Esener,
Email: gert.salentijn@wur.nl S. C.; Messmer, B. T. Nucleic Acids Res. 2016, 44, No. e96.

15594 https://dx.doi.org/10.1021/acs.analchem.0c03740
Anal. Chem. 2020, 92, 15587−15595
Analytical Chemistry pubs.acs.org/ac Article

(16) Anfossi, L.; Di Nardo, F.; Cavalera, S.; Giovannoli, C.; Baggiani,
C. Biosensors 2018, 9, 2.
(17) Oh, J.; Joung, H.-A.; Han, H. S.; Kim, J. K.; Kim, M.-G.
Theranostics 2018, 8, 3189−3197.
(18) Kim, T. H.; Hahn, Y. K.; Kim, M. S. Micromachines 2020, 11,
269.
(19) Tsai, T.-T.; Huang, T.-H.; Chen, C.-A.; Ho, N. Y.-J.; Chou, Y.-
J.; Chen, C.-F. Sci. Rep. 2018, 8, 17319.
(20) Hemmig, E.; Temiz, Y.; Gö kçe, O.; Lovchik, R. D.;
Delamarche, E. Anal. Chem. 2020, 92, 940−946.
(21) Shin, S.; Choi, M.; Shim, J.; Park, S. Sens. Actuators, B 2020,
304, 127408.
(22) Oh, Y. K.; Joung, H.-A.; Han, H. S.; Suk, H.-J.; Kim, M.-G.
Biosens. Bioelectron. 2014, 61, 285−289.
(23) Sathishkumar, N.; Toley, B. J. Sens. Actuators, B 2020, 324,
128756.
(24) Bremer, M. G. E. G.; Smits, N. G. E.; Haasnoot, W. Anal.
Bioanal. Chem. 2009, 395, 119−126.
(25) Ross, G. M. S.; Bremer, M.; Wichers, J. H.; van Amerongen, A.;
Nielen, M. W. F. Biosensors 2018, 8, 130.
(26) Ross, G. M. S.; Salentijn, G. IJ.; Nielen, M. W. F. Biosensors
2019, 9, 143.
(27) Marsh, J. T.; Jayasena, S.; Gaskin, F.; Baumert, J. L.; Johnson, P.
Food Chem. 2020, 313, 126019.
(28) Eriksson, E.; Lysell, J.; Larsson, H.; Cheung, K. Y.; Filippini, D.;
Mak, W. C. Research 2019, 2019, 8079561.
(29) Schneider, C. A.; Rasband, W. S.; Eliceiri, K. W. Nat. Methods
2012, 9, 671−675.
(30) Taylor, S. L.; Nordlee, J. A.; Niemann, L. M.; Lambrecht, D. M.
Anal. Bioanal. Chem. 2009, 395, 83−92.
(31) Qian, S.; Bau, H. H. Anal. Biochem. 2003, 322, 89−98.
(32) Ramos-Vara, J. A.; Miller, M. A. Vet. Pathol. 2014, 51, 42−87.
(33) Shire, S. J.; Shahrokh, Z.; Liu, J. J. Pharm. Sci. 2004, 93, 1390−
1402.
(34) Koets, M.; Renström, A.; Zahradnik, E.; Bogdanovic, J.;
Wouters, I. M.; van Amerongen, A. Journal of Environmental
Monitoring 2011, 13, 3475−3480.
(35) Duo, J.; Bruno, J.; Kozhich, A.; David-brown, D.; Luo, L.;
Kwok, S.; Santockyte, R.; Haulenbeek, J.; Liu, R.; Hamuro, L.;
Peterson, J. E.; Piccoli, S.; DeSilva, B.; Pillutla, R.; Zhang, Y. J.
Bioanalysis 2018, 10, 559−576.
(36) Duo, J.; Bruno, J.; Piccoli, S.; DeSilva, B.; Zhang, Y. J.
Bioanalysis 2017, 9, 193−207.
(37) Bray, D.; Lay, S. Proc. Natl. Acad. Sci. U.S.A. 1997, 94, 13493−
13498.

15595 https://dx.doi.org/10.1021/acs.analchem.0c03740
Anal. Chem. 2020, 92, 15587−15595

You might also like